Nitrogen-Enriched Hierarchically Porous Carbons Prepared from

Aug 19, 2014 - The sample HPC-800 activated at 800 °C possesses a high specific surface ... available by participants in CrossRef's Cited-by Linking ...
7 downloads 0 Views 6MB Size
Research Article www.acsami.org

Nitrogen-Enriched Hierarchically Porous Carbons Prepared from Polybenzoxazine for High-Performance Supercapacitors Liu Wan,†,‡ Jianlong Wang,† Lijing Xie,† Yahui Sun,†,‡ and Kaixi Li*,† †

Institute of Coal Chemistry, Chinese Academy of Sciences, Taiyuan 030001, P.R. China Graduate University of Chinese Academy of Sciences, Beijing 100049, P.R. China



S Supporting Information *

ABSTRACT: Nitrogen-enriched hierarchically porous carbons (HPCs) were synthesized from a novel nitrile-functionalized benzoxazine based on benzoxazine chemistry using a soft-templating method and a potassium hydroxide (KOH) chemical activation method and used as electrode materials for supercapacitors. The textural and chemical properties could be easily tuned by adding a soft template and changing the activation temperature. The introduction of the soft-templating agent (surfactant F127) resulted in the formation of mesopores, which facilitated fast ionic diffusion and reduced the internal resistance. The micropores of HPCs were extensively developed by KOH activation to provide large electrochemical double-layer capacitance. As the activation temperature increased from 600 to 800 °C, the specific surface area of nitrogen-enriched carbons increased dramatically, micropores were enlarged, and more meso/macropores were developed, but the nitrogen and oxygen content decreased, which affected the electrochemical performance. The sample HPC-800 activated at 800 °C possesses a high specific surface area (1555.4 m2 g−1), high oxygen (10.61 wt %) and nitrogen (3.64 wt %) contents, a hierarchical pore structure, a high graphitization degree, and good electrical conductivity. It shows great pseudocapacitance and the largest specific capacitance of 641.6 F g−1 at a current density of 1 A g−1 in a 6 mol L−1 KOH aqueous electrolyte when measured in a three-electrode system. Furthermore, the HPC-800 electrode exhibits excellent rate capability (443.0 F g−1 remained at 40 A g−1) and good cycling stability (94.3% capacitance retention over 5000 cycles). KEYWORDS: supercapacitor, polybenzoxazine, hierarchically porous carbon, nitrogen functionalities and chemical stability, and low cost.5,6 However, the intrinsically small micropores ( 0.9) denote the existence of macropores, which is consistent with the results of the SEM images. In a word, HPC600, HPC-700, and HPC-800 exhibit hierarchical porous

NPC-600 exhibits type I isotherms, indicating typical microporous carbon materials. Compared with HPC-0, no obvious meso/macropores can be found in NPC-600 from its PSD curve. This proves that the mesopores in HPC-0 were successfully produced by a soft-templating method, which arose from decomposition of surfactant F127 during the carbonization process. For samples HPC-600, HPC-700, and HPC-800, their isotherms all show type IV characteristics with 15587

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Table 2. Elemental and XPS Analysis and Electrochemical Properties of the Samples elemental analysis (wt %)

Cg (F g−1)

XPS analysis (atom %)

sample

C

H

N

O

C

N

O

HPC-0 NPC-600 HPC-600 HPC-700 HPC-800

78.35 79.28 80.96 82.23 85.10

2.32 2.25 1.47 0.87 0.65

6.26 4.82 5.19 4.55 3.64

13.07 13.65 12.38 12.35 10.61

81.21 83.35 84.10 85.35 87.23

5.89 4.27 4.88 4.16 3.27

12.90 12.38 11.02 10.49 9.50

1Ag

−1

40 A g−1

48.0 254.4 375.4 279.7 641.6

63.8 231.1 133.3 443.0

Figure 4. XPS spectra of all samples: (a) survey spectra; (b) N 1s; (c) O 1s; (d) nitrogen and oxygen species of the carbon materials.

structures: abundant micropores, limited mesopores, and a few macropores. Compared with NPC-600, the micropore size distribution of HPC-600 became wider as a result of the introduction of surfactant F127. This was probably caused by KOH corrosion with more accessible pore channels in the precursor HPC-0. Remarkably, the average mesopore size of HPC-600 maintained about 2.1 nm, revealing that the mesostructure of HPC-0 was not influenced by chemical activation. Besides, the activation temperature also has a remarkable impact on the pore structure. As the KOH

activation process deepened at higher temperature, new micropores were generated and the ultramicropores were widened, resulting in higher surface area and total pore volume with a lower fraction of the micropore volume for HPC-800 (see Table 1). The average mesopore size increased slightly from 2.1 to 2.3 nm with an increase of the activation temperature from 600 to 800 °C, indicating that the mesopores were enlarged by KOH etching under severe activation conditions. Such a trend for the mesopore structure evolution of the HPCs was further demonstrated by the PSDs obtained 15588

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Table 3. XPS Peak Positions and Relative Surface Concentrations of Oxygen and Nitrogen Species Obtained by Fitting the N 1s and O 1s Spectra sample

N-6 (398.5 eV)

N-5 (400.2 eV)

N-Q (401 eV)

N-X (403 eV)

O-I (531.3 eV)

O-II (532.5 eV)

O-III (535.0 eV)

HPC-0 NPC-600 HPC-600 HPC-700 HPC-800

21.84 33.38 35.02 33.31 30.77

58.10 42.95 50.66 42.82 43.40

14.41 13.82 12.11 13.85 15.00

5.66 9.85 2.21 10.02 10.83

23.78 25.27 18.52 40.48 16.02

73.88 70.18 79.97 58.07 81.73

2.34 4.54 1.51 1.45 2.25

electrical conductivity of the carbon-based materials, facilitate the accessibility of the electrolyte ions, and enhance the capacitance performance.18,19 The contents of the nitrogen and oxygen atoms in HPC-600−HPC-800 decreased evidently with an increase of the activation temperature because of the poor thermal stability of the nitrogen and oxygen species at high temperature. The chemical state of the nitrogen and oxygen species on the surface of the carbon materials was further studied by XPS analysis (Figure 4). Four types of nitrogen species can be distinguished on all sample surfaces: pyridinic nitrogen (N-6 at 398.5 eV), pyrrolic or pyridonic nitrogen (N-5 at 400.2 eV), quaternary nitrogen (N-Q at 401 eV), and oxidized nitrogen (N-X at 403 eV), as shown in Table 3 and Figure 4d.35 This reveals that the N-6 and N-5 species are predominant (more than 74% of the total nitrogen atoms) for all samples, which might be derived from the transformation of amide groups and triazine rings in PBZs after thermal treatment. Such high contents of N-6 and N-5 in PBZ-based nitrogen-containing porous carbons are much higher than many other nitrogen-rich carbon materials.17−19 More importantly, N-6 and N-5 are considered to be electrochemically active in an alkaline aqueous solution to provide main pseudocapacitance, which is significant for nitrogen-doped carbon materials to increase the capacitance.13 Besides, the presence of N-Q, which are inset into the carbon matrix and bonded to three carbon atoms, can effectively benefit electron transfer and improve the conductivity of carbonaceous materials.36 Thus, it can be inferred that HPC-800 with the highest N-Q content exhibits better conductivity than other samples. The conclusion was further proved by the electrical conductivity measurement results. The electrical conductivity σ values for HPC-0, NPC600, HPC-600, HPC-700, and HPC-800 are 2.86, 3.22, 3.24, 3.39, and 3.44, respectively, which are much higher than those of many activated carbons and ordered mesoporous carbons without incorporating any heteroatoms (0.2−2 S cm−1).2,5 This can be attributed to the local graphite-like microstructure and large amount of nitrogen and oxygen species in the carbon network. The electrical conductivity evidently enhances with an increase of the N-Q content (see Table 3). Moreover, an increase of the degree of graphitization also results in an increase of the conductivity of the nitrogen-enriched porous carbons. Thus, HPC-800, possessing the highest N-Q content and highest degree of graphitization, shows the best electrical conductivity among the five samples. Besides, with an increase of the activation temperature, the percentage of N-Q increased but N-X evidently decreased. This reveals that N-Q was more thermally stable than N-X and that N-5 and N-6 were partially converted into N-Q at high temperature. Furthermore, compared with NPC-600, the slightly higher nitrogen content in both bulk and surface for HPC-600 can be attributed to decomposition of the surfactant F127 together with the oxygen species in the precursor during thermal treatment.

by the BJH method (see Figure S1 in the Supporting Information, SI). Furthermore, macropores with size ranging from 60 to 100 nm for HPC-600, HPC-700, and HPC-800 were created by KOH corrosion and volatilization of unstable nitrogen and oxygen functionalities at high temperature. Because of PBZs’ high chemical and thermal stability, all obtained carbon materials exhibit high yield (more than 45%), as shown in Table 1. The graphitic property of the carbon samples is investigated by wide-angle XRD patterns and Raman spectra. The XRD patterns of NPC-600 and the HPCs are given in Figure 3c. Two typical, broad diffraction peaks are observed at around 2θ = 25 and 44°, which belong to diffraction of the (002) and (100) planes of the hexagonal graphitic carbon, respectively.31 This indicates that all of the nitrogen-containing porous carbons are amorphous and nongraphitized. HPC-0, NPC-600, and HPC600 prepared at 600 °C have similar, broad diffraction peaks. However, when the activation temperature increased from 600 to 800 °C, the intensity of the (100) peak was gradually enhanced, implying the improvement of the graphitization degree of HPC-800. Figure 3d shows the Raman spectra of all samples. Two strong peaks at 1350 cm−1 (D band) and 1595 cm−1 (G band) are observed for all carbon materials, respectively. In detail, the D band refers to the disordered and imperfect structures in the carbonaceous materials, and the G band is related to the vibration of a sp2-hybridized carbon in the graphite crystallites.32 The samples HPC-0, NPC-600, and HPC-600 obtained at 600 °C exhibit an intense D band, implying the existence of abundant defects, such as a large number of nitrogen and oxygen atoms. The intensity ratio of D and G bands (ID/IG) represents the graphitic degree of the carbon materials.33,34 The values of ID/IG are 1.04, 1.01, 0.97, 0.92, and 0.90 for HPC-0, NPC-600, HPC-600, HPC-700, and HPC-800, respectively. This reveals that the introduction of surfactant F127 or chemical activation resulted in small changes of the graphitic degree of the carbon materials because of the similar chemical compositions and structures of HPC-0, NPC600, and HPC-600. However, as the thermal treatment temperature increased from 600 to 800 °C, the vibration of the D band became weaker and ID/IG for HPC-600−HPC-800 decreased, indicating the higher degree of graphitization and a decrease of the disordered structure for HPC-800. This can be attributed to the loss of nitrogen and oxygen species at high activation temperature. These results are in good agreement with XRD analysis. The chemical compositions of the obtained carbon materials are shown in Table 2. It can be seen that all samples contain relatively high nitrogen species (3.64−6.26 wt %) and oxygen species (10.61−13.65 wt %). These heteroatoms can only be derived from the intrinsic nitrogen and oxygen components in the nitrile-functionalized PBZs (see Figure 1). More importantly, the nitrogen and oxygen atoms in the carbon matrix are considered to effectively improve the wettability and 15589

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Figure 5. Electrochemical performance of all HPC electrodes measured in a three-electrode system: CV curves for all samples at scan rates of (a) 5 and (b) 100 mV s−1. GCD curves for all samples at current densities of (c) 1 and (d) 10 A g−1. (e) Specific capacitance as a function of the discharge current density for all samples. (f) Nyquist plots in the frequency range of 10 mHz to 100 kHz.

6 M KOH aqueous electrolyte.35 The influence of the textural properties and surface chemistry of the HPCs on the capacitance was further estimated by electrochemical characterization. 3.2. Electrochemical Measurements. The electrochemical performance of all nitrogen-containing porous carbons as electrode materials for supercapacitors was evaluated by a three-electrode system in a 6 M KOH aqueous electrolyte. Figure 5a shows typical CV curves of all five electrodes at a scan rate of 5 mV s−1. It can be observed that all of them present a quasi-rectangular voltammogram shape at a low scan rate, exhibiting typical characteristics of EDLCs, indicative of them being excellent candidates as electrode materials.40 The nonactivated carbon material HPC-0 exhibits the smallest

The O 1s spectra of HPCs can be resolved into quinone (OI), phenolic hydroxyl or ether (O-II), and carboxyl (O-III) peaks centered at 531.3, 532.5, and 535.0 eV, respectively (Table 3 and Figure 4d). Among these three oxygen functional groups, quinone groups in the carbon matrix are not electrochemically active in the reversible redox reactions in an alkaline medium.37 Instead, reduction of O-II and deprotonation of O-III exhibited quasi-reversible pseudocapacitances.38,39 It is noticeable that all samples have high oxygen content (exceed 10 wt % according to the elemental analysis results) with extremely high O-II and O-III contents (exceeding 74 atom % of the total oxygen atoms), as shown in Table 3. Thus, the large amount of O-II, O-III, N-6, and, N-5 functionalities in the PBZ-based HPCs can generate great pseudocapacitance in a 15590

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Figure 6. Discharge time of the pseudocapacitance parts of (a) NPC-600, (b) HPC-600, (c) HPC-700, and (d) HPC-800 electrodes at 1 A g−1.

indicating that such a microporous carbon with narrow micropore size distribution is unfavorable for quick electrolyte ion diffusion. This confirms that the hierarchical pore structure is beneficial for fast ionic transportation and penetration within the macro/mesopore channels and adequate utilization of the interior surfaces through low-resistance pathways.43 Thus, HPC-600−HPC-800 show improved capacitive performance in comparison with the carbonized sample HPC-0 and microporous carbon NPC-600. Figure 5c clearly shows the GCD curves of all five electrodes at a current density of 1 A g−1. It can be seen that GCD curves of all five electrodes are nearly linear and symmetrical with slight curvature, indicating good capacitive properties and electrochemical reversibility. The deviation to linearity for all GCD curves is typical of a pseudocapacitance contribution, which confirms the presence of Faradaic capacitance. Compared with the other four electrodes, the discharge curve of the HPC-0 electrode exhibits the shortest discharge time, and its capacitance is as low as 48.0 F g−1 at 1 A g−1, which is consistent with the CV results. The NPC-600 electrode shows much shorter discharge time than the HPC-600 electrode. The presence of meso/macropores in HPC-600 not only makes the inner micropore surface more electrochemically accessible for electrolyte ions and more charges to be accumulated in the micropores but also facilitates the fast diffusion of electrolyte ions in the pore channels at high current densities. Moreover,

current density response compared with the KOH-activated carbons at the same scan rate, indicating the smallest specific capacitance of HPC-0. This can be attributed to its lowest specific surface area and thus a small electric double-layer (EDL) contribution. Noticeably, the CV curve for the HPC600 electrode shows small oxidation and reduction peaks in addition to rectangular shape, indicating the existence of pseudocapacitive reactions.41 This can be ascribed to its highest nitrogen content among the four KOH activated samples. However, other samples’ CV curves exhibit no obvious evidence of redox peaks but broadened pseudopeaks, indicating that the great contribution of EDLC as well as the pseudocapacitance. Although HPC-0 contains the largest number of nitrogen and oxygen species in the five samples, the low specific surface area and less developed porosity of HPC-0 make it difficult for fast charge transfer during the Faradaic reactions and result in poor capacitive behavior. When the scan rate increases from 5 to 100 mV s−1, the CV curves for the HPC-600−HPC-800 electrodes still remain a less rectangular shape (Figures 5b and S2 in the SI), implying good capacitive performance for quick charge/discharge operations. The broad pseudocapacitance peaks on the CV curves for HPC-600 ∼ HPC-800 electrodes can be ascribed to the combined contribution of the double-layer and Faradaic redox capacitances.42 However, the CV curve for the NPC-600 electrode becomes seriously distorted at high scan rates, 15591

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Figure 7. Electrochemical performance of the HPC-800 electrode measured in a two-electrode system. (a) CV curves at different scan rates. (b) GCD curves at different current densities.

Figure 5e exhibits the specific capacitance as a function of the discharge current density for all samples. It can be seen that the specific capacitance of HPC-0 drops from 48.0 to 12.6 F g−1 with a decrease in the current density from 1.0 to 10 A g−1, retaining only 26.2% of its initial specific capacitance. For the NPC-600, HPC-600, HPC-700, and HPC-800 electrodes, the values of the final specific capacitances at 40.0 A g−1 are 25.1, 61.6, 47.7, and 69.0% of the initial values at 1.0 A g−1, respectively (see Table 2). This reveals that HPC-600−HPC800 retain excellent capacitive behavior at high current densities, owing to their hierarchical porous structures and good electrical conductivity. However, the nonactivated carbon HPC-0 electrode and microporous carbon NPC-600 electrode show poor rate capability because of the less developed specific surface area and low porosity for HPC-0 and large diffusion resistance for NPC-600. With an increase of the activation temperature, both HPC600 and HPC-700 have similar pore structure and oxygen content, but the capacitance of the HPC-700 electrode with higher surface area dramatically decreases with a decrease of the nitrogen content at 1−40 A g−1. The result is also consistent with the much smaller pseudocapacitance obtained from TP for the HPC-700 electrode than that of the HPC-600 electrode. It can be further concluded that nitrogen functional groups play a more important role than oxygen functional groups in this work. Besides, compared with HPC-700, HPC-800 exhibits similar surface area, smaller micropore surface area and micropore volume, as well as lower bulk nitrogen content, but the HPC-800 electrode demonstrates the highest specific capacitances at all current densities (Figure 5e), resulting from its high specific surface area, an optimal PSD, abundant meso/ macropores with highly accessible surface for more nitrogen and oxygen functional groups to provide large pseudocapacitance, high degree of graphitization, and superior electrical conductivity. This reveals that, except for the high content of the N-6, N-5, O-II, and O-III functional groups in the porous carbon materials, the accessibility of the nitrogen-doped carbon surface for the electrolyte ions also plays an important role in improving the pseudocapacitance.45,46 Figure 5f exhibits Nyquist plots of all five electrodes. The HPC-0 electrode shows a small semicircle at the frequency region and an inclined line at the low-frequency region, indicating poor capacitive performance and high ion-diffusion

the higher contents of electrochemically active N-6, N-5, O-II, and O-III functionalities for HPC-600 generate more pseudocapacitance than that of the NPC-600 electrode. Therefore, the specific capacitance of the HPC-600 electrode increases by 47.6% at 1 A g−1 and 262.2% at 40 A g−1 compared to the NPC-600 electrode at the same current densities, respectively (see Table 2). Compared with the HPC-600 electrode, a decrease of the discharge time with an increase of the activation temperature for the HPC-700 electrode can be attributed to a decrease of significant pseudocapacitance, which results from the loss of electrochemically active nitrogen and oxygen species for HPC-700. Even when the current density increases to 20 A g−1 (Figure 5d), the GCD curves for all five electrodes still maintain a nearly linear-like shape with evident IR drop. In detail, there is a much larger IR drop of the NPC600 electrode than of the HPC-600 electrode at a high current density of 20 A g−1. This indicates that the HPC-600 electrode has less overall resistance because of its wider micropore distribution and the presence of meso/macropores. Besides, the IR drops for HPC-600−HPC-800 electrodes become smaller with an increase of the activation temperature because of the higher fractions of meso/macropores. In order to further investigate the contribution of the pseudocapacitance from nitrogen and oxygen functionalities to the total specific capacitance, the discharge curve can be divided into two parts because of the different extents of inflection at about −0.2 V in the GCD curve (Figure 6). The linear part of the time dependence of the potential (linear region) represents the EDLC. The other part of the discharge curve (curve region) is not strictly symmetrical and slightly distorted, which is considered to be the result of the combination of the EDLC and pseudocapacitance.44 The extended lines of the linear parts are drawn in Figure 6 to calculate the pseudocapacitance by nitrogen and oxygen functionalities. TD stands for the discharge time of the EDLC, and TP stands for the discharge time of the pseudocapacitance. Therefore, the pseudocapacitances calculated from TP of the NPC-600, HPC-600, HPC-700, and HPC800 electrodes are 136.3, 243.1, 113.6, and 317.2 F g−1 at 1 A g−1, which accounts for about 53.6, 64.8, 40.6, and 49.4% of the corresponding total capacitances, respectively. This confirms the pronounced contribution of pseudocapacitance for the nitrogen-enriched HPCs. 15592

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Figure 8. (a) Cycling stability of the HPC-800 electrode at a current density of 5 A g−1. (b) CV curves of the 1st and 5000th cycles of the HPC-800 electrode at a scan rate of 100 mV s−1 in a three-electrode system.

Table 4. Comparison of the Specific Capacitance Parameters of Different Nitrogen-Containing Porous Carbons in 6 M KOH in the Literature material nitrogen-doped porous carbons from gelatin hierarchical nitrogen-doped porous carbon from endothelium corneum gigeriae galli nitrogen-enriched carbon from melamine mica nitrogen-enriched carbon nanowires from polyaniline nanowires nitrogen-containing hydrothermal carbon nitrogen-doped ordered mesoporous carbon nitrogen-doped porous carbon nitrogen-doped porous nanofibers NPC-600 HPC-600 HPC-700 HPC-800

SBET (m2 g−1)

N%

3012 2149.9

0.88 2.04

6 641 571 537 128 562.51 954.8 1097.3 1526.5 1555.4

13.5 7.04 4.4 13.10 14.12 7.22 4.82 5.19 4.55 3.64

resistance of the nonactivated sample.47 The NPC-600, HPC600, HPC-700, and HPC-800 electrodes show a similar Nyquist plot shape, a depressed semicircle on the Z′ axis at the highfrequency region, and a nearly straight line at the low-frequency region, indicating good capacitive behavior of the chemically activated samples. With an increase of the activation temperature, the shortened and less gradual sloping line at high-tomedium frequency for HPC-600−HPC-800 electrodes implies lower diffusion resistance,48 which is attributed to the more developed meso/macropores at high activation temperature. On the basis of the same testing conditions, the diameter of the semicircle at high frequency refers to the charge-transfer resistance of the electrode materials, as shown in the inset of the expanded high-frequency region of Figure 5f.49 The chargetransfer resistances for the HPC-0, NPC-600, HPC-600, HPC700, and HPC-800 electrodes are 0.77, 0.93, 0.65, 0.61, and 0.38 Ω, respectively. This reveals that the HPC-800 electrode possesses the best ionic conductivity over the other samples because of its highest electrical conductivity and highest degree of graphitization. In addition, the transfer resistance of the HPC-600 electrode was much smaller than that of the NPC600 electrode. This confirms that the introduction of surfactant F127 can effectively tailor the pore structure and thus facilitate the transfer and penetration of the electrolyte ions into the inner pores.

Cg (F g−1) 385 198 198 327 220 227 224 202 254.4 375.4 279.7 641.6

Cg/SBET (F m−2)

current density (A g−1)

0.128 0.092

1.0 1.0

18 19

33 0.510 0.385 0.423 1.75 0.359 0.266 0.342 0.183 0.412

0.05 0.1 0.1 0.2 1.0 1.0 1.0 1.0 1.0 1.0

27 50 51 52 53 55 this this this this

ref

work work work work

To better study the electrochemical behavior of the HPC800 electrode as a real capacitor in a 6 M KOH solution electrolyte, a symmetrical two-electrode configuration was constructed. Figure 7a shows the CV curves of the HPC-800 electrode at different scan rates (5−200 mV s−1) using a twoelectrode system. This reveals that the current densities increase evidently with an increase of the scan rates, implying the good rate capacity. It is noticeable that there is a sharp peak at −0.9 V at a slow scan rate (5 mV s−1), which is related to the redox reactions. The CV curve of the HPC-800 electrode still maintains a rectangular-like shape at a scanning rate as high as 200 mV s−1, but the NPC-600, HPC-600, and HPC-700 electrodes exhibit distorted CV shapes due to increased electrical resistance (see Figure S3b in the SI). Moreover, the specific capacitances for a single HPC-800 electrode (563.9 and 347.9 F g−1 at 1 and 20 A g−1, respectively) calculated from the discharge curve in Figure 7b by using a two-electrode setup are slightly smaller than the values at the same current densities by using a three-electrode setup because of the different measurement methods (see Table S1 in the SI). It also demonstrates a good rate capability with a high capacitance retention ratio of 61.7% at 40 A g−1 for the HPC-800 electrode. Furthermore, the HPC-800 electrode exhibits good electric conductivity because its IR drop is minimal at a high current density of 10 A g−1 among the five HPC electrodes (see Figure 15593

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

of 69%, which is ascribed to its high surface area, optimal PSD, abundant nitrogen and oxygen functional groups attached to the carbon network, high graphitization degree, and electrical conductivity. It is worth noting that no obvious capacitance loss is observed over 5000 charge/discharge cycles, clearly demonstrating good cycling durability. The nitrogen-enriched HPC represents an alternative promising candidate for highperformance electrode materials for supercapacitors.

S3d in the SI), which is consistent with the electrical conductivity measurement results. In order to evaluate the cycling stability of the HPC-800 electrode, GCD studies were performed at a current density of 5.0 A g−1 between −0.9 and 0 V in the 6 M aqueous KOH electrolytes using a three-electrode system. After 5000 cycles, the specific capacitance of the HPC-800 electrode decreased from 530.2 to 500.0 F g−1, with a capacitance retention ratio of 94.3% (Figure 8a), indicative of excellent long-term cycling durability. The cycling durability of theHPC-800 electrode can be confirmed by the integral areas surrounded by the CV curves of the 1st and 5000th cycles at a scan rate of 100 mV s−1, as depicted in Figure 8b. Table 4 shows a comparison of the specific capacitances Cg and the interfacial capacitance Cs (Cs = Cg/SBET) of HPCs prepared from PBZs with other nitrogen-enriched porous carbons in 6 M KOH in the literature. The specific capacitances of the HPC-600 and HPC-700 electrodes in this work are comparable to those previously reported for the nitrogenenriched porous carbons (Table 4). Remarkably, the specific capacitance of the HPC-800 electrode is the highest at present compared with many nitrogen-containing porous carbon materials previously reported.18,19,50−53 Furthermore, according to the theoretical capacitance per area of 0.20 F m−2 for activated carbons as electrode materials,54 the theoretically calculated capacitance values for the NPC-600 and HPC-600− HPC-800 electrodes cannot exceed 191.0, 219.5, 305.3, and 311.1 F g−1, respectively. This indicates that, except for HPC700, at least 24.9, 41.5, and 51.5% of the total specific capacitances for the NPC-600, HPC-600, and HPC-800 electrodes were derived from the contribution of pseudocapacitance by the large number of nitrogen and oxygen functional groups in the carbon matrix, respectively. These estimated values are comparable to the calculated results obtained from the discharge time of the pseudocapacitance. In a word, the HPC-800 electrode shows extremely high specific capacitance and outstanding rate capability and cycling stability because of the following superior properties: (1) high specific surface area to provide enough surface sites to form EDL; (2) the optimal hierarchical pore structure with high fraction of meso/macropores, which provides more accessible micropore surface, reduces diffusion resistance, and facilitates fast ionic transportation in the pore channels; (3) the existence of high contents of electrochemically active nitrogen and oxygen functional groups, which not only improves the hydrophilicity and wettability of the HPC and its affinity with the electrolyte but also participates in reversible redox reactions and generates pronounced pseudocapacitance; (4) high graphitization degree and good electrical conductivity, which is also critical for high-performance capacitors.



ASSOCIATED CONTENT

S Supporting Information *

PSDs, electrochemical and electrochemical capacitive performances of all HPC electrodes, and specific capacitances of the samples. This material is available free of charge via the Internet at http://pubs.acs.org.



AUTHOR INFORMATION

Corresponding Author

*Tel.: 86 351 4250292. Fax: 86 351 4250292. E-mail: likx@ sxicc.ac.cn. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work is supported by the National Natural Science Foundation of China (No. 51002166, 51172251, and 51061130536), International Cooperation Project of the Ministry of Science and Technology (No. 2010DFB90690-4), International Cooperation Project of the Shanxi Province (No. 2013081016), the Shanxi Province Science and Technology Developing Project (20120313006-4), and Shanxi Province Science and Technology Industrialization Project (2013071051).



REFERENCES

(1) Winter, M.; Brodd, R. J. What Are Batteries, Fuel cells, and Supercapacitors? Chem. Rev. 2008, 104, 4245−4270. (2) Zhang, L. L.; Zhao, X. S. Carbon-Based Materials as Supercapacitor Electrodes. Chem. Soc. Rev. 2009, 38, 2520−2531. (3) Simon, P.; Gogotsi, Y. Materials for electrochemical capacitors. Electrochemical Capacitors for Energy Management. Nat. Mater. 2008, 7, 845−854. (4) Fic, K.; Meller, M.; Frackowiak, E. Strategies for Enhancing the Performance of Carbon/Carbon Supercapacitors in Aqueous Electrolytes. Electrochim. Acta 2014, 128, 210−217. (5) Zhai, Y.; Dou, Y.; Zhao, D.; Fulvio, P. F.; Mayes, R. T.; Dai, S. Carbon Materials for Chemical Capacitive Energy Storage. Adv. Mater. 2011, 23, 4828−4850. (6) Inagaki, M.; Konno, H.; Tanaike, O. Carbon Materials for Electrochemical Capacitors. J. Power Sources 2010, 195, 7880−7903. (7) Frackowiak, E.; Béguin, F. Carbon Materials for the Electrochemical Storage of Energy in Capacitors. Carbon 2001, 39, 937−950. (8) Nishihara, H.; Kyotani, T. Templated Nanocarbons for Energy Storage. Adv. Mater. 2012, 24, 4473−4498. (9) Xia, K.; Gao, Q.; Jiang, J.; Hu, J. Hierarchical Porous Carbons with Controlled Micropores and Mesopores for Supercapacitor Electrode Materials. Carbon 2008, 46, 1718−1726. (10) Chmiola, J.; Yushin, G.; Gogotsi, Y.; Portet, C.; Simon, P.; Taberna, P. L. Anomalous Increase in Carbon Capacitance at Pore Sizes Less Than 1 Nanometer. Science 2006, 313, 1760−1763. (11) Huang, J.; Sumpter, B. G.; Meunier, V. Theoretical Model for Nanoporous Carbon Supercapacitors. Angew. Chem., Int. Ed. 2008, 47, 520−524. (12) Yuan, C. Z.; Gao, B.; Shen, L. F.; Yang, D. S.; Hao, L.; Lu, X. J.; Zhang, F.; Zhang, L. J.; Zhang, X. G. Hierarchically Structured

4. CONCLUSIONS Novel nitrogen-enriched HPCs were for the first time successfully synthesized from a novel PBZ by a soft-templating method together with KOH activation. The introduction of a soft-templating agent and chemical activation temperature has a great influence on the textural properties and the surface chemistry of HPCs and thus influenced their capacitive performance. The sample HPC-800 electrode obtained by KOH activation at 800 °C exhibits a maximum specific capacitance of 641.6 F g−1 at a current density of 1 A g−1 in a 6 M KOH electrolyte. Moreover, its specific capacitance can still retain 443.0 F g−1 at 40 A g−1 with a capacitance retention ratio 15594

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

Carbon-Based Composites: Design, Synthesis and Their Application in Electrochemical Capacitors. Nanoscale 2011, 3, 529−545. (13) Ania, C. O.; Khomenko, V.; Raymund-Piñero, E.; Parra, J. B.; Béguin, F. The Large Electrochemical Capacitance of Microporous Doped Carbon Obtained by Using a Zeolite Template. Angew. Chem., Int. Ed. 2007, 17, 1828−1836. (14) Hall, P. J.; Mirzaeian, M.; Fletcher, S. I.; Sillars, F. B.; Rennie, A. J. R.; Shitta-Bey, G. O.; Wilson, G.; Cruden, A.; Carter, R. Energy Storage in Electrochemical Capacitors: Designing Functional Materials to Improve Performance. Energy Environ. Sci. 2010, 3, 1238−1251. (15) Hulicova-Jurcakova, D.; Puziy, A. M.; Poddubnaya, O. I.; Suárez-García, F.; Tascón, J. M. D.; Lu, G. Q. Highly Stable Performance of Supercapacitors from Phosphorus-Enriched Carbons. J. Am. Chem. Soc. 2009, 131, 5026−5027. (16) Kim, C.; Ngoc, B. T. N.; Yang, K. S.; Kojima, M.; Kim, Y. A.; Kim, Y. J.; Endo, M.; Yang, S. C. Self-Sustained Thin Webs Consisting of Porous Carbon Nanofibers for Supercapacitors via the Electrospinning of Polyacrylonitrile Solutions Containing Zinc Chloride. Adv. Mater. 2007, 19, 2341−2346. (17) Yan, J.; Wei, T.; Qiao, W.; Fan, Z.; Zhang, L.; Li, T.; Zhao, Q. A High-Performance Carbon Derived from Polyaniline for Supercapacitors. Electrochem. Commun. 2010, 12, 1279−1282. (18) Xu, B.; Hou, S.; Cao, G.; Wu, F.; Yang, Y. S. Sustainable Nitrogen-Doped Porous Carbon with High Surface Areas Prepared from Gelatin for Supercapacitors. J. Mater. Chem. 2012, 22, 19088− 19093. (19) Hong, X.; Hui, K. S.; Zeng, Z.; Hui, K. N.; Zhang, L.; Mo, M.; Li, M. Hierarchical Nitrogen-Doped Porous Carbon with High Surface area Derived from Endothelium Corneum Gigeriae Galli for HighPerformance Supercapacitor. Electrochim. Acta 2014, 130, 464−469. (20) Yagci, Y.; Kiskan, B.; Ghosh, N. N. Recent Advancement on PolybenzoxazineA Newly Developed High Performance Thermoset. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 5565−5576. (21) Ghosh, N. N.; Kiskan, B.; Yagci, Y. PolybenzoxazinesNew High Performance Thermosetting Resins: Synthesis and Properties. Prog. Polym. Sci. 2007, 32, 1344−1391. (22) Liu, X.; Gu, Y. Study on the Volumetric Expansion of Benzoxazine Curing with Different Catalysts. J. Appl. Polym. Sci. 2002, 84, 1107−1113. (23) Agag, T.; Arza, C. R.; Maurer, F. H. J.; Ishida, H. Primary Amine-Functional Benzoxazine Monomers and Their Use for AmideContaining Monomeric Benzoxazines. Macromolecules 2010, 43, 2748−2758. (24) Sawaryn, C.; Landfester, K.; Taden, A. Benzoxazine Miniemulsions Stabilized with Polymerizable Nonionic Benzoxazine Surfactants. Macromolecules 2010, 43, 8933−8941. (25) Sawaryn, C.; Landfester, K.; Taden, A. Benzoxazine Miniemulsions Stabilized with Multifunctional Main-chain Benzoxazine Protective Colloids. Macromolecules 2011, 44, 5650−5658. (26) Hao, G. P.; Li, W. C.; Qian, D.; Wang, G. H.; Zhang, W. P.; Zhang, T.; Wang, A. Q.; Schüth, F.; Bongard, H. J.; Lu, A. H. Structurally Designed Synthesis of Mechanically Stable Poly(Benzoxazine-co-Resol)-Based Porous Carbon Monoliths and Their Application as High-Performance CO2 Capture Sorbents. J. Am. Chem. Soc. 2011,, 133,, 11378-−11388.. (27) Cao, G. P.; Chen, W. J.; Liu, X. B. Synthesis and Thermal Properties of the Thermosetting Resin based on Cyano Functionalized Benzoxazine. Polym. Degrad. Stab. 2008, 93, 739−744. (28) Meng, Y.; Gu, D.; Zhang, F.; Shi, Y.; Yang, H.; Li, Z.; Yu, C.; Tu, B.; Zhao, D. Ordered Mesoporous Polymers and Homologous Carbon Frameworks: Amphiphilic Surfactant Templating and Direct Transformation. Angew. Chem. 2005, 117, 7215−7221. (29) Baqar, M.; Agag, T.; Huang, R.; Maia, J.; Qutubuddin, S.; Ishida, H. Mechanistic Pathways for the Polymerization of MethylolFunctional Benzoxazine Monomers. Macromolecules 2012, 45, 8119− 8125. (30) Brunovska, Z.; Lvon, R.; Ishida, H. Thermal Properties of Phthalonitrile Functional Polybenzoxazines. Thermochim. Acta 2000, 357−358, 195−203.

(31) Zhao, Q.; Wang, X.; Wu, C.; Liu, J.; Wang, H.; Gao, J.; Zhang, Y. Supercapacitive Performance of Hierarchical Porous Carbon Microspheres Prepared by Simple One-Pot Method. J. Power Sources 2014, 254, 10−17. (32) Jin, H.; Wang, X.; Gu, Z.; Polin, J. Carbon Materials from High Ash Biochar for Supercapacitor and Improvement of Capacitance with HNO3 Surface Oxidation. J. Power Sources 2013, 236, 285−292. (33) Yuan, D.; Chen, J.; Zeng, J.; Tan, S. Preparation of Monodisperse Carbon Nanospheres for Electrochemical Capacitors. Electrochem. Commun. 2008, 10, 1067−1070. (34) Wang, Q.; Yan, J.; Wang, Y.; Ning, G.; Fan, Z.; Wei, T.; Cheng, J.; Zhang, M.; Jing, X. Template Synthesis of Hollow Carbon Spheres Anchored on Carbon Nanotubes for High Rate Performance Supercapacitors. Carbon 2013, 52, 209−218. (35) Hulicova-Jurcakova, D.; Kodama, M.; Shirashi, S.; Hatori, H.; Zhu, Z. H.; Lu, G. Q. Nitrogen-Enriched Nonporous Carbon Electrodes with Extraordinary Supercapacitance. Adv. Funct. Mater. 2009, 19, 1800−1809. (36) Wu, J.; Zhang, D.; Wang, Y.; Hou, B. R. Electrocatalytic Activity of Nitrogen-Doped Graphene Synthesized via a One-Pot Hydrothermal Process Towards Oxygen Reduction Reaction. J. Power Sources 2013, 227, 185−190. (37) Andreas, H. A.; Conway, B. E. Examination of the Double-Layer Capacitance of an High Specific-Area C-Cloth Electrode as Titrated from Acidic to Alkaline pHs. Electrochim. Acta 2006, 51, 6510−6520. (38) Kerisit, S.; Schwenzer, B.; Vijayakumar, M. Effects of OxygenContaining Functional Groups on Supercapacitor Performance. J. Phys. Chem. Lett. 2014, 5, 2330−2334. (39) Fujisawa, K.; Cruz-Silva, R.; Yang, K. S.; Kim, Y. A.; Hayashi, T.; Endo, M.; Terrones, M.; Dresselhaus, M. S. Importance of Open, Heteroatom-Decorated Edges in Chemically Doped-Graphene for Supercapacitor Applications. J. Mater. Chem. A 2014, 2, 9532−9540. (40) Wickramaratne, N. P.; Xu, J.; Wang, M.; Zhu, L.; Dai, L.; Jaroniec, M. Nitrogen Enriched Porous Carbon Spheres: Attractive Materials for Supercapacitor Electrodes and CO2 Adsorption. Chem. Mater. 2014, 26, 2820−2828. (41) Candelaria, S. L.; Garcia, B. B.; Liu, D.; Cao, G. Nitrogen Modification of Highly Porous Carbon for Improved Supercapacitor Performance. J. Mater. Chem. 2012, 22, 9884−9889. (42) Lee, Y. H.; Chang, K. H.; Hu, C. C. Differentiate the Pseudocapacitance and Double-Layer Capacitance Contributions for Nitrogen-Doped Reduced Graphene Oxide in Acidic and Alkaline Electrolytes. J. Power Sources 2013, 227, 300−308. (43) Guo, Y.; Shi, Z.; Chen, M. M.; Wang, C. Y. Hierarchical Porous Carbon Derived from Sulfonated Pitch for Electrical Double Layer Capacitors. J. Power Sources 2014, 252, 235−43. (44) Yang, X.; Wu, D.; Chen, X.; Fu, R. Nitrogen-Enriched Nanocarbons with a 3-D Continuous Mesopore Structure from Polyacrylonitrile for Supercapacitor Application. J. Phys. Chem. C 2010, 114, 8581−8586. (45) Wang, D. W.; Li, F.; Chen, Z. G.; Gao, Q. L.; Cheng, H. M. Synthesis and Electrochemical Property of Boron-Doped Mesoporous Carbon in Supercapacitor. Chem. Mater. 2008, 20, 7195−7200. (46) Chmiola, J.; Yushin, G.; Dash, R.; Gogotsi, Y. Effect of Pore Size and Surface Area of Carbide Derived Carbons on Specific Capacitance. J. Power Sources 2006, 158, 765−772. (47) Yu, P.; Li, Y.; Zhao, X.; Wu, L.; Zhang, Q. Graphene-Wrapped Polyaniline Nanowire Arrays on Nitrogen-Doped Carbon Fabric as Novel Flexible Hybrid Electrode Materials for High-Performance Supercapacitor. Langmuir 2014, 30, 5306−5313. (48) Nasini, U. B.; Bairi, V. G.; Ramasahayam, S. K.; Bourdo, S. E.; Viswanathan, T.; Shaikh, A. U. Phosphorous and Nitrogen Dual Heteroatom Doped Mesoporous Carbon Synthesized via Microwave Method for Supercapacitor Application. J. Power Sources 2014, 250, 257−265. (49) Xu, B.; Hou, S. S.; Zhang, F. L.; Cao, G. P.; Chu, M.; Yang, Y. S. Nitrogen-Doped Mesoporous Carbon Derived from Biopolymer as Electrode Material for Supercapacitors. J. Electroanal. Chem. 2014, 712, 146−150. 15595

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596

ACS Applied Materials & Interfaces

Research Article

(50) Yuan, D.; Zhou, T.; Zhou, S.; Zou, W.; Mo, S.; Xia, N. NitrogenEnriched Carbon Nanowires from the Direct Carbonization of Polyaniline Nanowires and its Electrochemical Properties. Electrochem. Commun. 2011, 13, 242−246. (51) Zhao, L.; Fan, L. Z.; Zhou, M. Q.; Guan, H.; Qiao, S.; Antonietti, M.; Titirici, M. M. Nitrogen-Containing Hydrothermal Carbons with Superior Performance in Supercapacitors. Adv. Mater. 2010, 22, 5202−5206. (52) Wei, J.; Zhou, D.; Sun, Z.; Deng, Y.; Xia, Y.; Zhao, D. A Controllable Synthesis of Rich Nitrogen-Doped Ordered Mesoporous Carbon for CO2 Capture and Supercapacitors. Adv. Funct. Mater. 2013, 23, 2322−2328. (53) Chen, X. Y.; Chen, C.; Zhang, Z. J.; Xie, D. H.; Deng, X.; Liu, J. W. Nitrogen-Doped Porous Carbon for Supercapacitor with LongTerm Electrochemical Stability. J. Power Sources 2013, 230, 50−58. (54) Frackowiak, E. Carbon Materials for Supercapacitor Application. Phys. Chem. Chem. Phys. 2007, 9, 1774−1785. (55) Chen, L. F.; Zhang, X. D.; Liang, H. W.; Kong, M. G.; Guan, Q. F.; Chen, P.; Wu, Z. Y.; Yu, S. H. Synthesis of Nitrogen-Doped Porous Carbon Nanofibers as an Efficient Electrode Material for Supercapacitors. ACS Nano 2012, 6, 7092−7102.

15596

dx.doi.org/10.1021/am504564q | ACS Appl. Mater. Interfaces 2014, 6, 15583−15596